Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Article

Discord in Concordance Cosmology and Anomalously Massive


Early Galaxies
Stacy McGaugh 1

1 Department of Astronomy, Case Western Reserve University; stacy.mcgaugh@case.edu

Abstract: Cosmological parameters are constrained by a wide variety of observations. We examine


the concordance diagram for modern measurements of the Hubble constant, the shape parameter
from large scale structure, the cluster baryon fraction, and the age of the universe, all from non-CMB
data. There is good agreement for H0 = 73.24 ± 0.38 km s−1 Mpc−1 and Ωm = 0.237 ± 0.015. This
concordance value is indistinguishable from the WMAP3 cosmology but is not consistent with that
arXiv:2312.03127v2 [astro-ph.CO] 17 Jan 2024

of Planck: there is a tension in Ωm as well as H0 . These tensions have emerged as progressively


higher multipoles have been incorporated into CMB fits. This temporal evolution is suggestive of
a systematic effect in the analysis of CMB data at fine angular scales, and may be related to the
observation of unexpectedly massive galaxies at high redshift. These are overabundant relative to
ΛCDM predictions by an order of magnitude at z > 7. Such massive objects are anomalous and
could cause gravitational lensing of the surface of last scattering in excess of the standard calculation
made in CMB fits, potentially skewing the best-fit cosmological parameters and contributing to the
Hubble tension.

Keywords: Background radiation, cosmic; Early universe; Galaxy formation; Gravitational lensing;
Hubble constant

1. Concordance Cosmology
The concordance cosmology (ΛCDM) emerged in the 1990s thanks to observational
advances that constrained a broad variety of cosmological parameters [1,2]. These led to the
surprising recognition that the mass density was less than the critical density [3], in conflict
with then-standard SCDM cosmology and Inflationary theory. Retaining the latter implied
a second surprise in the form of the cosmological constant such that Ωm + ΩΛ = 1, which
in turn predicted that the expansion rate of the universe would currently be accelerating
[4]. This unlikely-seeming prediction was subsequently observed in the Hubble diagram
of Type Ia SN [5,6]. This picture was further corroborated by the location of the first
peak of the acoustic power spectrum of the CMB being consistent with a flat (Ωk = 0)
Citation: McGaugh, S. Cosmic Robertson-Walker geometry [7,8].
Discord. Preprints 2022, 1, 0. The range of concordance parameters narrowed with the completion of the Hubble
https://doi.org/ Space Telescope Key Project to Measure the Hubble Constant [9], and a ‘vanilla’ set of
Publisher’s Note: MDPI stays neutral ΛCDM parameters emerged with1 h = 0.7 and Ωm = 0.3 [10] some twenty years ago.
with regard to jurisdictional claims in Since that time, a tension [11–13] has emerged between direct measurements of the Hubble
published maps and institutional affil- constant [9,14–20] and that obtained from fits to the acoustic power spectrum of the cosmic
iations. microwave background obtained by the Planck mission [21]. This tension was not present
in earlier data from WMAP [22]. Indeed, one of the most persuasive arguments in favor
of vanilla ΛCDM was the concordance of WMAP cosmological parameters with other
observations, especially H0 . Now that a tension has emerged, it is natural to ask when and
Copyright: © 2022 by the author. where things went amiss.
Licensee MDPI, Basel, Switzerland.
To this end, it is interesting to update the Ωm -H0 concordance diagram that was
This article is an open access article
pivotal in establishing ΛCDM in the first place [2]. Figure 1 shows modern values of the
distributed under the terms and
same constraints used in Ref. [2] nearly three decades ago. These include the SH0ES direct
conditions of the Creative Commons
measurement of the Hubble constant H0 = 73.04 ± 1.04 km s−1 Mpc−1 [19], the shape
Attribution (CC BY) license (https://
creativecommons.org/licenses/by/
4.0/).
1 h = H0 /(100 km s−1 Mpc−1 ).
2 of 13

parameter Ωm h = 0.168 ± 0.016 from large scale structure [23], the baryon fraction of
clusters of galaxies Ωm h1/2 = 0.221 ± 0.031 [24], and the age of the globular cluster M92,
13.80 ± 0.75 Gyr [25]. For the age constraint, we assume a flat geometry as an open universe
without a cosmological constant would be too young to have a concordance region. For
simplicity, we also assume the age of the universe is indistinguishable from the age of the
globular cluster, which presumably formed very early, and surely within the first 750 Myr
(the uncertainty in its age).

H0
WMAP3
CM
B

e
Ag Planck
LSS
Cluster baryon fraction

Figure 1. The concordance region (white space) in Ωm -H0 space [2] where the allowed regions
(colored bands) of many constraints intersect. Illustrated constraints include a direct measurement
of the Hubble constant [19, black], the age of a flat universe containing the globular cluster M92
[25, light blue], the cluster baryon fraction [24, dark blue], and the shape parameter from large scale
stucture [23, orange]. This modern rendition of the concordance region in consistent with a WMAP3
cosmology [orange point, 22] but not with that of Planck [yellow point, 21]. The green line represents
the covariance of CMB fits, Ωm h3 = 0.09633 ± 0.00030 [21].

The data in Fig. 1 tell very much the same story as it has for the entirety of this
century. Taking these observations at face value, we recover a concordance cosmology
with Ωm = 0.235 ± 0.015 and h = 0.7304 ± 0.0104. This is completely consistent with the
WMAP3 [22] cosmology (Ωm = 0.241 ± 0.034, h = 0.732 ± 0.032), but less so with later
results.

2. Variations on the Concordance Diagram


The most accurate constraint in Fig. 1 is the direct measurement of the Hubble con-
stant from SH0ES [19], so it is worth considering what happens if we adopt other val-
ues. There are several groups that have independently achieved ∼ 1% precision with
random errors < 1 km s−1 Mpc−1 [14,19,20]. Considering only the random error, the
CosmicFlows-4 measurement H0 = 74.6 ± 0.8 [20] is consistent with the SH0ES value
(H0 = 73.04 ± 1.04 km s−1 Mpc−1 ), differing by only 1.2 σ, while the CCHP value of
H0 = 69.8 ± 0.8 km s−1 Mpc−1 [14] is marginally inconsistent at the ∼ 2.5 σ level.
Taking these measurements of H0 and their statistical uncertainties at face value, we
can reconstruct the concordance diagram for each. Combining the CCHP H0 = 69.8 ±
0.8 km s−1 Mpc−1 [14] with the other constraints leads to a concordance region with Ωm =
0.247 ± 0.017. Doing the same for the CosmicFlows-4 H0 = 74.6 ± 0.8 [20] gives Ωm =
3 of 13

0.233 ± 0.013. Treating these as the practical range of measured Hubble constants, the
corresponding range of mass density is 0.22 ≤ Ωm ≤ 0.264. This range is consistent with
local estimates of the mass density [26,27] and with WMAP cosmologies but not with the
Ωm = 0.315 ± 0.007 of the Planck cosmology [21].

H0

WMAP3 CM WMAP3 CM
B B

H0

e
Ag

Ag
Planck Planck
LSS LSS
Cluster baryon fraction Cluster baryon fraction

Figure 2. The same as Fig. 1 but a lower (left) and higher (right) choice of H0 to illustrate the range of
systematic uncertainty in direct determinations of the Hubble constant. Neither choice reconciles the
concordance of local cosmological constraints with the Planck cosmology.

Fig. 3 summarizes the status of the Hubble tension as of this writing. Local mea-
surements range from 69 to 75 km s−1 Mpc−1 obtained by a variety of methods and cali-
brations. This subject is the focus of much research with frequent updates to essentially
the same result, e.g., Freedman et al. [14] and Freedman [28], so we restrict Fig. 3 to a
single entry from each distinct group. We highlight the most precise measurements with
statistical uncertainties ≤ 3 km s−1 Mpc−1 . The weighted mean of these values gives
H0 = 73.24 ± 0.38 km s−1 Mpc−1 . The range of mass density consistent with this and the
other constraints in the concordance diagram is Ωm = 0.237 ± 0.015. This precise estimate
from the concordance of local observations is indistinguishable from the WMAP3 [22]
cosmology.b
The above assessment is both straightforward and naive, as it ignores systematic
errors, which are difficult to quantify. The history of the distance scale predisposes us to
suspect a systematic error somewhere in the calibration of the distance ladder. This view
is encouraged by the apparent tension between Cepheid [19] and TRGB [14] calibrations.
However, there is no tension between these two calibrators in other analyses [17,18,20], so
this may not be a difference between Cepheid and TRGB methods so much as it is between
the distinct TRGB methods employed by different groups [29,30].
An assessment of systematics is built into the SH0ES error bar [19], so by their as-
sessment, the tension with Planck remains highly significant (5σ). Independent assess-
ments using the Tully-Fisher relation [31] consistently find H0 consistent with the SH0ES
value or even a tad higher [12,20], which goes in the wrong direction to achieve consis-
tency with Planck. Indeed, application of the baryonic Tully-Fisher relation [32] excludes
H0 < 70.5 km s−1 Mpc−1 at 95% c.l. [17]; the asymmetric probability distribution skews
to higher rather than lower values, so H0 < 68 km s−1 Mpc−1 is practically excluded
barring a systematic error in the calibration of both Cepheids and the TRGB method.
The plausible amplitude of such a systematic calibration uncertainty is suggested to be
±1.3 km s−1 Mpc−1 by the CCHP [33]. This almost reconciles the low estimate of H0 from
CCHP with Planck: the tension is only 1.6 σ if both random and systematic uncertainties
are at their maximum and combine to go in the same direction. Setting aside the eternal
question of how much to believe the calibrations and the assessments of systematic un-
certainty [12,13], there are also geometric methods [15,16] that are independent of stellar
calibrators. These also favor H0 in the low to mid-70s (Fig. 3). Consequently, the essential
4 of 13

answer has not changed since the completion of the Hubble Space Telescope Key Project to
Measure the Hubble Constant [9]. Progress has been made in accuracy, which has improved
to the point that it is difficult to sustain the presumption that there must be a systematic
error in the distance scale [12].

Figure 3. Recent direct measurements of H0 [11] over time (left panel) including the earlier HST key
project [9, far left]. Measurements with statistical uncertainties ≤ 3 km s−1 Mpc−1 from independent
groups are color coded by calibrator: geometry [black: 15,16], Cepheids [blue: 9,19], TRGB [orange: 14],
or a combination [cyan: 17,18,20]. Statistical and systematic uncertainties are shown as dark and light
error bars, respectively; these are omitted for less accurate data (small open circles). Determinations
of H0 from WMAP [22,34–36, orange] and Planck [21,37, yellow] covary with the mass density as
Ωm h3 = 0.09633 ± 0.00030 [green line, 21], and have decreased steadily over time (right panel). For
comparison, the red square shows the latest CMB value in the left panel and the mean of the accurate
direct determinations in the right panel (at arbitrary Ωm ). The concordance region from Fig. 1 is
shown as the open region in the right panel; alternate concordance regions from Fig. 2 are shown as
light grey. The remaining region (dark grey) is excluded by those data.

Tension in ΛCDM is not restricted to the Hubble constant; it also appears in the
mass density. Indeed, it appears using modern measurements of the same observational
constraints that established the concordance cosmology in the first place [2]. A major
reason for the persistence of ΛCDM was the excellent agreement between early WMAP
cosmologies and the wealth of previous constraints that all indicated the same region of
concordance in Fig. 1.
A similar success does not extend to Planck cosmologies. This is true for any choice
of locally measured Hubble constant. While obvious for the higher estimates of H0 , even
the lowest value from the CCHP is not really in agreement with Planck so much as it is
not in strong tension with it. However, if we also consider the mass density implied in
Fig. 2, then the tension is magnified: not only is H0 too big, but Ωm is too low. Given the
different dependencies on h of the various constraints, there is no prospect of regaining
concordance. Even if a local measurement gave exactly the same Hubble constant as
Planck, the concordance region would remain at lower Ωm ≈ 0.27 rather than 0.315, simply
transferring the tension in H0 to one in Ωm . Indeed, the concordance window for any
H0 < 72 km s−1 Mpc−1 fails to contain the locus of Ωm h3 demanded by CMB data (the
green line in Fig. 3); there is no FLRW universe that satisfies all the illustrated constraints.
Faced with the choice between abandoning the FLRW cosmology and disbelieving
some aspect of the data, most people choose the latter. However, it is becoming difficult to
do so, as other tensions do exist. For example, there is a persistent tension in the power
spectrum normalization σ8 [38–40]. There is also a significant but oft-neglected tension in
the baryon density between deuterium [41] and lithium [42]. The deuterium value agrees
5 of 13

with CMB fits, so the general perception seems to be that lithium must be in systematic2
error. This is equivalent to assuming the Planck estimate of H0 is correct and all local
measurements must inevitably be in error.
The most consequential constraint considered here, after direct measurements of
H0 , is that on the product Ωm h of large scale structure from the 2dF Galaxy Redshift
Survey [45]. The 2dF value, Ωm h = 0.168 ± 0.016 [23], limits how large a mass density is
allowed, and precludes concordance with the higher value of Planck (Ωm h = 0.212 ± 0.005).
Ironically, the preliminary result of the same survey initially gave a more consistent result,
Ωm h = 0.20 ± 0.03 [46], but the answer changed as the larger survey volume was analyzed
[23]. The difference between the initial and final result of the 2dF survey gives cause to
ponder the extent to which the structure of the local universe is consistent with ΛCDM
[47–49], with some even calling into question the basic cosmological premise of isotropy
[50–54].
At this juncture, there are so many measurements that it is possible to find indications
of tensions in any cosmological parameter. The trick is to assess the credibility of each
while avoiding confirmation bias. It is also important to bear in mind the possibility that
the concordance window is indeed closed, and there is no viable FLRW universe. It is
conceivable that the strange parameters of ΛCDM are merely the best FLRW approximation
to some deeper theory [55–60].

3. Covariance and the Hubble Tension


The best fit to the Planck CMB data, H0 = 67.36 ± 0.54 km s−1 Mpc−1 [21], clearly
differs from the bulk of the accurate “local” determinations [9,14–20]. The weighted mean
of these values differs from the Planck value by nearly 9σ, so formally the tension is
real. Though systematic uncertainties are inevitable at some level, that level is around
±1.3 km s−1 Mpc−1 [33], not the ∼ 6 km s−1 Mpc−1 that would be required to reconcile the
two: the tension remains 4.5 σ for plausible systematic uncertainties. In other words, as
the precision of local distance scale measurements have approached 1%, their accuracy has
reached ∼ 2%, but we need a systematic of 9% to make the problem go away.
Local measurements of the distance scale and fits to the CMB measure completely
different things. The local measurements are all some realization of the traditional Hubble
program in which distances and redshifts to nearby galaxies are measured, and the slope of
the velocity–distance relation is obtained. This is an empirical procedure. In contrast, fits to
the power spectrum of the CMB at z = 1090 are model (ΛCDM) dependent, and the results
are sensitive to all of the cosmic parameters simultaneously with inevitable covariance.
The strongest covariances affecting the Hubble parameter is with the mass density. These
covary in CMB fits approximately as Ωm h3 = 0.09633 ± 0.00030 [21].
The covariance of H0 with Ωm is apparent in Fig. 3. There is also a correlation
with time: the best-fit combination has marched steadily along the trench of constant
Ωm h3 from WMAP3 with H0 = 73.2 ± 3.2 km s−1 Mpc−1 [22] to Planck with H0 =
67.36 ± 0.54 km s−1 Mpc−1 [21]. It is the CMB result that has progressively deviated
from the concordance ΛCDM region established around the turn of the millennium, not
local measurements.
The temporal progression of the best-fit values of H0 and Ωm is a flag that suggests
something systematic in the CMB analyses rather than traditional Hubble constant deter-
minations. The latter scatter approximately as expected: the rms variation of the accurate
data in Fig. 3 is 1.78 km s−1 Mpc−1 while the size of the statistical uncertainties anticipate
that it should be 1.55 km s−1 Mpc−1 . This implies that systematic uncertainties are not
contributing much to the scatter, though one must always be cautious of a calibration issue
that would translate all results without affecting the scatter [33].

2 There was no discrepancy prior to the appearance of CMB constraints [43]. Since then, lithium estimates have
remain unchanged while deuterium changed suddenly to come into concordance [44].
6 of 13

The CMB data from independent experiments are in good agreement where they
overlap, so the temporal progression is not a measurement error. What has changed over
time is the availability of data on ever finer scales, tracing the power spectrum to higher
multipoles ℓ. Something subtle seems to be tipping the covariance of parameters to favor
lower H0 along the trench of constant Ωm h3 as higher ℓ have been incorporated into the
fits.

4. Massive Galaxies at High Redshift


An important recent development is the observation of unexpectedly massive galaxies
at high redshift. This may seem unrelated at first, but CMB data have reached the level of
precision where the subtle effects of gravitational lensing by masses intervening between
ourselves and the surface of last scattering cannot be ignored. If massive galaxies assembled
anomalously early, then gravitational lensing by these objects may have a systematic impact
on CMB fits.

Figure 4. The number density of galaxies as a function of their stellar mass, color coded by redshift,
as labeled. The left panel shows predicted stellar mass functions [61, lines] with the corresponding
data [62, circles]. The right panel shows the ratio of the observed to predicted density of galaxies.
There is a clear excess of massive galaxies at high redshifts.

Galaxies are expected to assemble gradually through the hierarchical merger of many
progenitors in ΛCDM. This process takes a long time. A typical massive galaxy is predicted
to have assembled half its stars around z ≈ 0.7 when the universe is ∼ 7 Gyr old, roughly
half its current age [63,64]. In contrast, there are many galaxies that are observed to be
massive and already quiescent at z ≈ 3 [65–68] when the universe is only ∼ 2 Gyr old.
These appear to have formed half of their stars in the first gigayear (z ≳ 6), and are
consistent with the traditional picture of a monolithic early type galaxy that formed early
(z ≥ 10) and followed an approximately exponential star formation history with a short
(∼ 1 Gyr) e-folding time [66]. This comes as a surprise3 to the predictions of hierarchical
structure formation in ΛCDM [72,73].
Though highlighted by recent JWST observations at z ≈ 10 [74–79], the discrepancy is
already apparent by z ≈ 3 in earlier observations [65,66]. Figure 4 illustrates the situation
shortly before the launch of JWST. The observed luminosity function of galaxies broadly
resembles that predicted, but there is a systematic excess of bright, massive galaxies. This
persists over all redshifts z > 3 and holds in all environments, both in clusters and in the
field [66].

3 The formation of massive galaxies at z ≥ 10 was predicted in advance by Bob Sanders in the context of MOND
[69], in which case the expansion history of the early universe may differ from ΛCDM [70]. The problems
posed by massive early galaxies may also be relieved if the universe is much older than generally thought [71].
7 of 13

Figure 5. The number density of galaxies as a function of their rest-frame ultraviolet absolute
magnitude observed by JWST, a proxy for stellar mass at high redshift. The left panel shows predicted
luminosity functions [80, lines], color coded by redshift, as labeled. Data in the corresponding redshift
bins are shown as squares [81], circles [82], and triangles [83]. The right panel shows the ratio of
the observed to predicted density of galaxies. The observed luminosity function barely evolves, in
contrast to the prediction of substantial evolution as the first dark matter halos assemble. There is a
large excess of bright galaxies at the highest redshifts observed.

The ‘impossibly early galaxy’ problem [65] has become more severe with new obser-
vations from JWST [84], extending to z > 9 (Fig. 5). Over this redshift range, the luminosity
function is predicted to evolve rapidly [80]. There is a good physical reason for this, as
this is the epoch of hierarchical mass assembly in ΛCDM when fragmentary protogalaxies
should first be forming before subsequently assembling into more massive galaxies at much
later times. There is not sufficient time in ΛCDM to assemble the observed mass into single
objects [72,73], hence the ragged luminosity functions in the highest redshift4 bins.
These JWST results are new, and will no doubt be debated for some time. Valid
concerns can be raised over the veracity of photometric redshifts and the relation of
ultraviolet starlight to dark matter halo mass. However, the anomalous presence of early
massive galaxies found by JWST (Fig. 5) corroborates previous work (Fig. 4) for which
spectroscopic redshifts are available and for which the assessment of stellar mass is more
robust. It is therefore hard to avoid the conclusion that galaxies grew too big too fast.

5. Gravitational Lensing of the CMB


The presence of massive galaxies in the early universe is an anomaly to the ΛCDM
structure formation paradigm. They should not be there [72,73]. The existence of such
galaxies violates the assumptions that underpin fits to the acoustic power spectrum of the
CMB, so may lead to systematic errors in the assessment of cosmological parameters based
on those fits. In particular, the calculation of gravitational lensing would be impacted [85].
Gravitational lensing of the surface of last scattering by intervening masses is an effect
that becomes important at high ℓ, beginning to impact the fit for ℓ > 800. Intriguingly,
restricting the fit of the Planck data to ℓ < 800 returns a higher Hubble constant [21]
consistent with that found by WMAP [36]. Consequently, concordance with local data is
maintained for low-ℓ data; the discrepancy has only appeared as high-ℓ data have been
incorporated into the fits. So perhaps there is a systematic effect due to the gravitational
lensing of the surface of last scattering.

4 A luminosity density of ∼ 7 × 10−6 mag.−1 Mpc−3 is estimated at z ≈ 16 [81]. This point is omitted from Fig. 5
because the corresponding prediction is not a number: galaxies bright enough to observe should not yet exist.
8 of 13

Figure 6. Top: unbinned Planck data [86] with the best-fit power spectrum (left) and a model with
H0 = 73 km s−1 Mpc−1 (right) with ΩCDM scaled to keep Ωm h3 = 0.09633 constant. Bottom: the ratio
of binned data and model to the best fit. The darker line is the model including the expected effects
of lensing; the lighter line is the power spectrum emergent from the surface of last scattering before
lensing. If there is an anomalous source of lensing from massive early galaxies then the mapping from
emergent to observed power spectrum will be miscalculated and the best-fit cosmological parameters
may be in error.

We have seen above that massive galaxies appear anomalously early in the history
of the universe. This implies a stronger lensing effect than expected. In order to compute
the effects of lensing on the observed CMB, lensing potentials are extrapolated forward in
time by assuming linear growth of the perturbations observed in the CMB. The JWST data
imply that this assumption is violated, but it is hardwired into the calculations used to fit
CMB data. This implies that the computation of lensing underestimates the real effect by
assuming a growth rate that is less than observed.
High mass galaxies at high redshift are anomalous, so we do not know how to calculate
the lensing potentials they represent. We can, however, use CAMB [87] to compute the
CMB power spectrum to see how the predicted power varies in the absence of this effect.
We do this in Fig. 6, which compares the Planck best fit with a model that has H0 =
73 km s−1 Mpc−1 . We scale the mass density of this model to maintain consistency with
the constraint Ωm h3 = 0.09633 [21]. We hold the mass density of baryons and neutrinos
constant, scaling that of cold dark matter to ΩCDM h2 = 0.109 for a total mass density of
Ωm = 0.248, consistent with the concordance region in Fig. 1. This gives an acoustic power
spectrum that is only distinguishable from the best-fit version upon close scrutiny (Fig. 6),
at a level where the effect of gravitational lensing is perceptible.
9 of 13

The lensing calculation, though subtle, plays an important role in the fit. Lensing blurs
the surface of last scattering, suppressing some of the oscillations that were emergent at
the time of recombination, as illustrated by the lighter lines in the lower panels of Fig. 6. If
the predicted amount of lensing is wrong because of the anomalously early appearance of
massive galaxies, the resulting fit will be systematically in error.
The power spectrum of the model with H0 = 73 km s−1 Mpc−1 is offset and tilted
from the Planck best fit. The offset at ℓ < 800 is mostly caused by covariance with the
optical depth to the surface of last scattering, which shifts the amplitude. WMAP found
τ = 0.089 ± 0.014 [36], rather higher than the best fit of Planck, τ = 0.0544 ± 0.0073 [21].
This is about what it takes to shift the model into agreement with the data at low ℓ, which
is part of why WMAP found a higher best-fit optical depth. There are, of course, other
covariances, especially with the amplitude of the power spectrum, σ8 , which itself is in
some tension with local measurements [38–40].
In addition to blurring the surface of last scattering, gravitational lensing also adds a
source of small scale power to the observed fluctuations that is not present in the power
spectrum emergent from the surface of last scattering. This will add poorly correlated
power at high ℓ. If there is an excess of such power from anomalously early galaxies, it
will tilt the power spectrum in the sense seen in Fig. 6. That is, if one starts with a model
like that with H0 = 73 km s−1 Mpc−1 , power will be added at high ℓ beyond what can be
explained by the model. If this happens but we are unaware of it, the fitting procedure will
indulge the covariance of all the cosmological parameters until a fit is found. Driving H0
down and Ωm up goes in the right direction to do this. It is therefore conceivable that the
Planck best-fit H0 is in systematic error due to anomalous gravitational lensing.

6. Conclusions
We have made an updated assessment of the concordance diagram that gave rise
to ΛCDM [2]. Using modern data, we find H0 = 73.24 ± 0.38 km s−1 Mpc−1 and Ωm =
0.237 ± 0.015. These values are essentially unchanged from twenty years ago, and are in
good agreement with the WMAP3 cosmology. They are not consistent with the Planck
cosmology: the tension is present in the mass density as well as the Hubble constant.
Cosmological parameters obtained from fits to the acoustic power spectrum of the
cosmic microwave background have gradually moved away from the concordance region
specified by independent constraints, moving steadily along the trench of constant Ωm h3
to lower H0 and higher Ωm leading to the current tension with the locally measured
value of the Hubble constant. This migration correlates with the inclusion of higher
multipoles in the fits, for which the effects of gravitational lensing are important. The
appearance of anomalously massive galaxies in the early universe as indicated by JWST
(and other) observations suggests that the impact of gravitational lensing on the CMB
may be underestimated. It is conceivable that this is a contributing factor to the temporal
migration of the best-fit cosmological parameters and the resulting Hubble tension. If so, it
may be the CMB value of H0 that is systematically in error rather than local determinations.
Massive galaxies at high redshift are anomalous in ΛCDM and would require new
physics to explain. We have refrained from speculating on what such new physics might
be, but note that it was predicted in advance by MOND [43,69]. In addition to specifying a
particular hypothesis for the new physics, it would further be necessary to incorporate the
predicted growth rate into a Boltzmann code in order to calculate the lensing effect on the
surface of last scattering. Such a calculation is beyond the scope of this work.

Funding: This research received no external funding.


Data Availability Statement: Data and models are available at the author’s websites:
astroweb.case.edu/ssm/data and astroweb.case.edu/ssm/models.
Conflicts of Interest: The author declares no conflict of interest.
10 of 13

Abbreviations
The following abbreviations are used in this manuscript:

CCHP Carnegie-Chicago Hubble Program


CMB Cosmic Microwave Background
JWST James Webb Space Telescope
ΛCDM Lambda Cold Dark Matter
SCDM Standard Cold Dark Matter
SH0ES Supernovae and H0 for the Equation of State of dark energy
WMAP Wilkinson Microwave Anisotropy Probe

References
1. Efstathiou, G.; Sutherland, W.J.; Maddox, S.J. The cosmological constant and cold dark matter. Nature 1990, 348, 705–707.
https://doi.org/10.1038/348705a0.
2. Ostriker, J.P.; Steinhardt, P.J. The observational case for a low-density Universe with a non-zero cosmological constant. Nature
1995, 377, 600–602. https://doi.org/10.1038/377600a0.
3. Davis, M.; Efstathiou, G.; Frenk, C.S.; White, S.D.M. The end of cold dark matter? Nature 1992, 356, 489–494. https:
//doi.org/10.1038/356489a0.
4. McGaugh, S.S. Cosmological constant. Nature 1996, 381, 483. https://doi.org/10.1038/381483b0.
5. Riess, A.G.; Filippenko, A.V.; Challis, P.; Clocchiatti, A.; Diercks, A.; Garnavich, P.M.; Gilliland, R.L.; Hogan, C.J.; Jha, S.; Kirshner,
R.P.; et al. Observational Evidence from Supernovae for an Accelerating Universe and a Cosmological Constant. AJ 1998,
116, 1009–1038, [arXiv:astro-ph/astro-ph/9805201]. https://doi.org/10.1086/300499.
6. Perlmutter, S.; Aldering, G.; Goldhaber, G.; Knop, R.A.; Nugent, P.; Castro, P.G.; Deustua, S.; Fabbro, S.; Goobar, A.; Groom, D.E.;
et al. Measurements of Ω and Λ from 42 High-Redshift Supernovae. ApJ 1999, 517, 565–586, [arXiv:astro-ph/astro-ph/9812133].
https://doi.org/10.1086/307221.
7. de Bernardis, P.; Ade, P.A.R.; Bock, J.J.; Bond, J.R.; Borrill, J.; Boscaleri, A.; Coble, K.; Crill, B.P.; De Gasperis, G.; Farese, P.C.;
et al. A flat Universe from high-resolution maps of the cosmic microwave background radiation. Nature 2000, 404, 955–959,
[arXiv:astro-ph/astro-ph/0004404]. https://doi.org/10.1038/35010035.
8. Page, L.; Nolta, M.R.; Barnes, C.; Bennett, C.L.; Halpern, M.; Hinshaw, G.; Jarosik, N.; Kogut, A.; Limon, M.; Meyer, S.S.; et al.
First-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Interpretation of the TT and TE Angular Power
Spectrum Peaks. ApJS 2003, 148, 233–241, [arXiv:astro-ph/astro-ph/0302220]. https://doi.org/10.1086/377224.
9. Freedman, W.L.; Madore, B.F.; Gibson, B.K.; Ferrarese, L.; Kelson, D.D.; Sakai, S.; Mould, J.R.; Kennicutt, Jr., R.C.; Ford, H.C.;
Graham, J.A.; et al. Final Results from the Hubble Space Telescope Key Project to Measure the Hubble Constant. ApJ 2001,
553, 47–72, [arXiv:astro-ph/0012376]. https://doi.org/10.1086/320638.
10. Tegmark, M.; Strauss, M.A.; Blanton, M.R.; Abazajian, K.; Dodelson, S.; Sandvik, H.; Wang, X.; Weinberg, D.H.; Zehavi, I.; Bahcall,
N.A.; et al. Cosmological parameters from SDSS and WMAP. Phys. Rev. D 2004, 69, 103501, [arXiv:astro-ph/astro-ph/0310723].
https://doi.org/10.1103/PhysRevD.69.103501.
11. Di Valentino, E.; Mena, O.; Pan, S.; Visinelli, L.; Yang, W.; Melchiorri, A.; Mota, D.F.; Riess, A.G.; Silk, J. In the realm of
the Hubble tension-a review of solutions. Classical and Quantum Gravity 2021, 38, 153001, [arXiv:astro-ph.CO/2103.01183].
https://doi.org/10.1088/1361-6382/ac086d.
12. Tully, R.B. The Hubble Constant: A Historical Review. arXiv e-prints 2023, p. arXiv:2305.11950, [arXiv:astro-ph.CO/2305.11950].
https://doi.org/10.48550/arXiv.2305.11950.
13. Cervantes-Cota, J.L.; Galindo-Uribarri, S.; Smoot, G.F. The Unsettled Number: Hubble’s Tension. Universe 2023, 9, 501,
[arXiv:physics.hist-ph/2311.07552]. https://doi.org/10.3390/universe9120501.
14. Freedman, W.L.; Madore, B.F.; Hatt, D.; Hoyt, T.J.; Jang, I.S.; Beaton, R.L.; Burns, C.R.; Lee, M.G.; Monson, A.J.; Neeley, J.R.; et al.
The Carnegie-Chicago Hubble Program. VIII. An Independent Determination of the Hubble Constant Based on the Tip of the
Red Giant Branch. ApJ 2019, 882, 34, [arXiv:astro-ph.CO/1907.05922]. https://doi.org/10.3847/1538-4357/ab2f73.
15. Wong, K.C.; Suyu, S.H.; Chen, G.C.F.; Rusu, C.E.; Millon, M.; Sluse, D.; Bonvin, V.; Fassnacht, C.D.; Taubenberger, S.; Auger, M.W.;
et al. H0LiCOW - XIII. A 2.4 per cent measurement of H0 from lensed quasars: 5.3σ tension between early- and late-Universe
probes. MNRAS 2020, 498, 1420–1439, [arXiv:astro-ph.CO/1907.04869]. https://doi.org/10.1093/mnras/stz3094.
16. Pesce, D.W.; Braatz, J.A.; Reid, M.J.; Riess, A.G.; Scolnic, D.; Condon, J.J.; Gao, F.; Henkel, C.; Impellizzeri, C.M.V.; Kuo, C.Y.;
et al. The Megamaser Cosmology Project. XIII. Combined Hubble Constant Constraints. ApJ 2020, 891, L1, [arXiv:astro-
ph.CO/2001.09213]. https://doi.org/10.3847/2041-8213/ab75f0.
17. Schombert, J.; McGaugh, S.; Lelli, F. Using the Baryonic Tully-Fisher Relation to Measure Ho . AJ 2020, 160, 71, [arXiv:astro-
ph.CO/2006.08615]. https://doi.org/10.3847/1538-3881/ab9d88.
18. Blakeslee, J.P.; Jensen, J.B.; Ma, C.P.; Milne, P.A.; Greene, J.E. The Hubble Constant from Infrared Surface Brightness Fluctuation
Distances. ApJ 2021, 911, 65, [arXiv:astro-ph.CO/2101.02221]. https://doi.org/10.3847/1538-4357/abe86a.
11 of 13

19. Riess, A.G.; Yuan, W.; Macri, L.M.; Scolnic, D.; Brout, D.; Casertano, S.; Jones, D.O.; Murakami, Y.; Anand, G.S.; Breuval, L.; et al. A
Comprehensive Measurement of the Local Value of the Hubble Constant with 1 km s−1 Mpc−1 Uncertainty from the Hubble Space
Telescope and the SH0ES Team. ApJ 2022, 934, L7, [arXiv:astro-ph.CO/2112.04510]. https://doi.org/10.3847/2041-8213/ac5c5b.
20. Tully, R.B.; Kourkchi, E.; Courtois, H.M.; Anand, G.S.; Blakeslee, J.P.; Brout, D.; Jaeger, T.d.; Dupuy, A.; Guinet, D.; Howlett, C.;
et al. Cosmicflows-4. ApJ 2023, 944, 94, [arXiv:astro-ph.CO/2209.11238]. https://doi.org/10.3847/1538-4357/ac94d8.
21. Planck Collaboration.; Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.; Barreiro,
R.B.; Bartolo, N.; et al. Planck 2018 results. VI. Cosmological parameters. A&A 2020, 641, A6, [arXiv:astro-ph.CO/1807.06209].
https://doi.org/10.1051/0004-6361/201833910.
22. Spergel, D.N.; Bean, R.; Doré, O.; Nolta, M.R.; Bennett, C.L.; Dunkley, J.; Hinshaw, G.; Jarosik, N.; Komatsu, E.; Page, L.; et al.
Three-Year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Implications for Cosmology. ApJS 2007, 170, 377–408,
[arXiv:astro-ph/astro-ph/0603449]. https://doi.org/10.1086/513700.
23. Cole, S.; Percival, W.J.; Peacock, J.A.; Norberg, P.; Baugh, C.M.; Frenk, C.S.; Baldry, I.; Bland-Hawthorn, J.; Bridges, T.; Cannon, R.;
et al. The 2dF Galaxy Redshift Survey: power-spectrum analysis of the final data set and cosmological implications. MNRAS
2005, 362, 505–534, [arXiv:astro-ph/astro-ph/0501174]. https://doi.org/10.1111/j.1365-2966.2005.09318.x.
24. Mantz, A.B.; Morris, R.G.; Allen, S.W.; Canning, R.E.A.; Baumont, L.; Benson, B.; Bleem, L.E.; Ehlert, S.R.; Floyd, B.; Herbonnet,
R.; et al. Cosmological constraints from gas mass fractions of massive, relaxed galaxy clusters. MNRAS 2022, 510, 131–145,
[arXiv:astro-ph.CO/2111.09343]. https://doi.org/10.1093/mnras/stab3390.
25. Ying, J.M.; Chaboyer, B.; Boudreaux, E.M.; Slaughter, C.; Boylan-Kolchin, M.; Weisz, D. The Absolute Age of M92. AJ 2023,
166, 18, [arXiv:astro-ph.SR/2306.02180]. https://doi.org/10.3847/1538-3881/acd9b1.
26. Mohayaee, R.; Tully, R.B. The Cosmological Mean Density and Its Local Variations Probed by Peculiar Velocities. ApJ 2005,
635, L113–L116, [arXiv:astro-ph/astro-ph/0509313]. https://doi.org/10.1086/499774.
27. Shaya, E.J.; Tully, R.B.; Hoffman, Y.; Pomarède, D. Action Dynamics of the Local Supercluster. ApJ 2017, 850, 207, [arXiv:astro-
ph.CO/1710.08935]. https://doi.org/10.3847/1538-4357/aa9525.
28. Freedman, W.L. Measurements of the Hubble Constant: Tensions in Perspective. ApJ 2021, 919, 16, [arXiv:astro-
ph.CO/2106.15656]. https://doi.org/10.3847/1538-4357/ac0e95.
29. Jang, I.S.; Hoyt, T.J.; Beaton, R.L.; Freedman, W.L.; Madore, B.F.; Lee, M.G.; Neeley, J.R.; Monson, A.J.; Rich, J.A.; Seibert, M. The
Carnegie-Chicago Hubble Program. IX. Calibration of the Tip of the Red Giant Branch Method in the Megamaser Host Galaxy,
NGC 4258 (M106). ApJ 2021, 906, 125, [arXiv:astro-ph.GA/2008.04181]. https://doi.org/10.3847/1538-4357/abc8e9.
30. Anand, G.S.; Tully, R.B.; Rizzi, L.; Riess, A.G.; Yuan, W. Comparing Tip of the Red Giant Branch Distance Scales: An Independent
Reduction of the Carnegie-Chicago Hubble Program and the Value of the Hubble Constant. ApJ 2022, 932, 15, [arXiv:astro-
ph.CO/2108.00007]. https://doi.org/10.3847/1538-4357/ac68df.
31. Tully, R.B.; Fisher, J.R. A new method of determining distances to galaxies. A&A 1977, 54, 661–673.
32. McGaugh, S.S.; Schombert, J.M.; Bothun, G.D.; de Blok, W.J.G. The Baryonic Tully-Fisher Relation. ApJ 2000, 533, L99–L102,
[arXiv:astro-ph/0003001]. https://doi.org/10.1086/312628.
33. Uddin, S.A.; Burns, C.R.; Phillips, M.M.; Suntzeff, N.B.; Freedman, W.L.; Brown, P.J.; Morrell, N.; Hamuy, M.; Krisciunas, K.;
Wang, L.; et al. Carnegie Supernova Project-I and -II: Measurements of H0 using Cepheid, TRGB, and SBF Distance Calibration to
Type Ia Supernovae. arXiv e-prints 2023, p. arXiv:2308.01875, [arXiv:astro-ph.CO/2308.01875]. https://doi.org/10.48550/arXiv.
2308.01875.
34. Komatsu, E.; Dunkley, J.; Nolta, M.R.; Bennett, C.L.; Gold, B.; Hinshaw, G.; Jarosik, N.; Larson, D.; Limon, M.; Page, L.; et al.
Five-Year Wilkinson Microwave Anisotropy Probe Observations: Cosmological Interpretation. ApJ Suppl. Ser. 2009, 180, 330–376.
https://doi.org/10.1088/0067-0049/180/2/330.
35. Komatsu, E.; Smith, K.M.; Dunkley, J.; Bennett, C.L.; Gold, B.; Hinshaw, G.; Jarosik, N.; Larson, D.; Nolta, M.R.; Page, L.; et al.
Seven-year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Cosmological Interpretation. ApJS 2011, 192, 18,
[arXiv:astro-ph.CO/1001.4538]. https://doi.org/10.1088/0067-0049/192/2/18.
36. Hinshaw, G.; Larson, D.; Komatsu, E.; Spergel, D.N.; Bennett, C.L.; Dunkley, J.; Nolta, M.R.; Halpern, M.; Hill, R.S.; Odegard, N.;
et al. Nine-year Wilkinson Microwave Anisotropy Probe (WMAP) Observations: Cosmological Parameter Results. ApJS 2013,
208, 19, [arXiv:astro-ph.CO/1212.5226]. https://doi.org/10.1088/0067-0049/208/2/19.
37. Planck Collaboration.; Ade, P.A.R.; Aghanim, N.; Armitage-Caplan, C.; Arnaud, M.; Ashdown, M.; Atrio-Barandela, F.; Aumont,
J.; Baccigalupi, C.; Banday, A.J.; et al. Planck 2013 results. XVI. Cosmological parameters. A&A 2014, 571, A16, [arXiv:astro-
ph.CO/1303.5076]. https://doi.org/10.1051/0004-6361/201321591.
38. Battye, R.A.; Charnock, T.; Moss, A. Tension between the power spectrum of density perturbations measured on large and small
scales. Phys. Rev. D 2015, 91, 103508, [arXiv:astro-ph.CO/1409.2769]. https://doi.org/10.1103/PhysRevD.91.103508.
39. Abbott, T.M.C.; Aguena, M.; Alarcon, A.; Allam, S.; Alves, O.; Amon, A.; Andrade-Oliveira, F.; Annis, J.; Avila, S.; Bacon, D.;
et al. Dark Energy Survey Year 3 results: Cosmological constraints from galaxy clustering and weak lensing. Phys. Rev. D 2022,
105, 023520, [arXiv:astro-ph.CO/2105.13549]. https://doi.org/10.1103/PhysRevD.105.023520.
40. Lange, J.U.; Hearin, A.P.; Leauthaud, A.; van den Bosch, F.C.; Xhakaj, E.; Guo, H.; Wechsler, R.H.; DeRose, J. Constraints
on S8 from a full-scale and full-shape analysis of redshift-space clustering and galaxy-galaxy lensing in BOSS. MNRAS 2023,
520, 5373–5393, [arXiv:astro-ph.CO/2301.08692]. https://doi.org/10.1093/mnras/stad473.
12 of 13

41. Hsyu, T.; Cooke, R.J.; Prochaska, J.X.; Bolte, M. The PHLEK Survey: A New Determination of the Primordial Helium Abundance.
ApJ 2020, 896, 77, [arXiv:astro-ph.GA/2005.12290]. https://doi.org/10.3847/1538-4357/ab91af.
42. Cyburt, R.H.; Fields, B.D.; Olive, K.A. An update on the big bang nucleosynthesis prediction for 7 Li: the problem worsens. JCAP
2008, 2008, 012, [arXiv:astro-ph/0808.2818]. https://doi.org/10.1088/1475-7516/2008/11/012.
43. McGaugh, S.S. A tale of two paradigms: the mutual incommensurability of ΛCDM and MOND. Canadian Journal of Physics 2015,
93, 250–259, [arXiv:astro-ph.CO/1404.7525]. https://doi.org/10.1139/cjp-2014-0203.
44. McGaugh, S.S. Confrontation of Modified Newtonian Dynamics Predictions with Wilkinson Microwave Anisotropy Probe First
Year Data. ApJ 2004, 611, 26–39, [arXiv:astro-ph/astro-ph/0312570]. https://doi.org/10.1086/421895.
45. Colless, M.; Dalton, G.; Maddox, S.; Sutherland, W.; Norberg, P.; Cole, S.; Bland-Hawthorn, J.; Bridges, T.; Cannon, R.; Collins, C.;
et al. The 2dF Galaxy Redshift Survey: spectra and redshifts. MNRAS 2001, 328, 1039–1063, [arXiv:astro-ph/astro-ph/0106498].
https://doi.org/10.1046/j.1365-8711.2001.04902.x.
46. Percival, W.J.; Baugh, C.M.; Bland-Hawthorn, J.; Bridges, T.; Cannon, R.; Cole, S.; Colless, M.; Collins, C.; Couch, W.; Dalton, G.;
et al. The 2dF Galaxy Redshift Survey: the power spectrum and the matter content of the Universe. MNRAS 2001, 327, 1297–1306,
[arXiv:astro-ph/astro-ph/0105252]. https://doi.org/10.1046/j.1365-8711.2001.04827.x.
47. Peebles, P.J.E.; Nusser, A. Nearby galaxies as pointers to a better theory of cosmic evolution. Nature 2010, 465, 565–569,
[arXiv:astro-ph.CO/1001.1484]. https://doi.org/10.1038/nature09101.
48. Peebles, P.J.E. Formation of the large nearby galaxies. MNRAS 2020, 498, 4386–4395, [arXiv:astro-ph.GA/2005.07588]. https:
//doi.org/10.1093/mnras/staa2649.
49. Neuzil, M.K.; Mansfield, P.; Kravtsov, A.V. The Sheet of Giants: Unusual properties of the Milky Way’s immediate neighbourhood.
MNRAS 2020, 494, 2600–2617, [arXiv:astro-ph.GA/1912.04307]. https://doi.org/10.1093/mnras/staa898.
50. Colin, J.; Mohayaee, R.; Rameez, M.; Sarkar, S. Evidence for anisotropy of cosmic acceleration. A&A 2019, 631, L13, [arXiv:astro-
ph.CO/1808.04597]. https://doi.org/10.1051/0004-6361/201936373.
51. Secrest, N.J.; von Hausegger, S.; Rameez, M.; Mohayaee, R.; Sarkar, S.; Colin, J. A Test of the Cosmological Principle with Quasars.
ApJ 2021, 908, L51, [arXiv:astro-ph.CO/2009.14826]. https://doi.org/10.3847/2041-8213/abdd40.
52. Secrest, N.J.; von Hausegger, S.; Rameez, M.; Mohayaee, R.; Sarkar, S. A Challenge to the Standard Cosmological Model. ApJ
2022, 937, L31, [arXiv:astro-ph.CO/2206.05624]. https://doi.org/10.3847/2041-8213/ac88c0.
53. Domènech, G.; Mohayaee, R.; Patil, S.P.; Sarkar, S. Galaxy number-count dipole and superhorizon fluctuations. JCAP 2022,
2022, 019, [arXiv:astro-ph.CO/2207.01569]. https://doi.org/10.1088/1475-7516/2022/10/019.
54. Jones, J.; Copi, C.J.; Starkman, G.D.; Akrami, Y. The Universe is not statistically isotropic. arXiv e-prints 2023, p. arXiv:2310.12859,
[arXiv:astro-ph.CO/2310.12859]. https://doi.org/10.48550/arXiv.2310.12859.
55. Penrose, R. Before the Big Bang: AN Outrageous New Perspective and its Implications for Particle Physics. In Proceedings of the
Proceedings of EPAC 2006, 2006, pp. 2759–2762.
56. Clifton, T.; Ferreira, P.G.; Padilla, A.; Skordis, C. Modified gravity and cosmology. Physics Reports 2012, 513, 1–189, [arXiv:astro-
ph.CO/1106.2476]. https://doi.org/10.1016/j.physrep.2012.01.001.
57. Penrose, R. On the Gravitization of Quantum Mechanics 1: Quantum State Reduction. Foundations of Physics 2014, 44, 557–575.
https://doi.org/10.1007/s10701-013-9770-0.
58. Penrose, R. On the Gravitization of Quantum Mechanics 2: Conformal Cyclic Cosmology. Foundations of Physics 2014, 44, 873–890.
https://doi.org/10.1007/s10701-013-9763-z.
59. Skordis, C.; Złośnik, T. New Relativistic Theory for Modified Newtonian Dynamics. Phys. Rev. Letters 2021, 127, 161302,
[arXiv:astro-ph.CO/2007.00082]. https://doi.org/10.1103/PhysRevLett.127.161302.
60. Nesbet, R.K. Conformal theory of gravitation and cosmic expansion. arXiv e-prints 2023, p. arXiv:2308.10399, [arXiv:gr-
qc/2308.10399]. https://doi.org/10.48550/arXiv.2308.10399.
61. Yung, L.Y.A.; Somerville, R.S.; Popping, G.; Finkelstein, S.L.; Ferguson, H.C.; Davé, R. Semi-analytic forecasts for JWST - II.
Physical properties and scaling relations for galaxies at z = 4-10. MNRAS 2019, 490, 2855–2879, [arXiv:astro-ph.GA/1901.05964].
https://doi.org/10.1093/mnras/stz2755.
62. Stefanon, M.; Bouwens, R.J.; Labbé, I.; Illingworth, G.D.; Gonzalez, V.; Oesch, P.A. Galaxy Stellar Mass Functions from z 10 to z 6
using the Deepest Spitzer/Infrared Array Camera Data: No Significant Evolution in the Stellar-to-halo Mass Ratio of Galaxies in
the First Gigayear of Cosmic Time. ApJ 2021, 922, 29, [arXiv:astro-ph.GA/2103.16571]. https://doi.org/10.3847/1538-4357/ac1
bb6.
63. Vogelsberger, M.; Genel, S.; Springel, V.; Torrey, P.; Sijacki, D.; Xu, D.; Snyder, G.; Bird, S.; Nelson, D.; Hernquist, L. Properties
of galaxies reproduced by a hydrodynamic simulation. Nature 2014, 509, 177–182, [arXiv:astro-ph.CO/1405.1418]. https:
//doi.org/10.1038/nature13316.
64. Henriques, B.M.B.; White, S.D.M.; Thomas, P.A.; Angulo, R.; Guo, Q.; Lemson, G.; Springel, V.; Overzier, R. Galaxy formation in
the Planck cosmology - I. Matching the observed evolution of star formation rates, colours and stellar masses. MNRAS 2015,
451, 2663–2680, [arXiv:astro-ph.GA/1410.0365]. https://doi.org/10.1093/mnras/stv705.
65. Steinhardt, C.L.; Capak, P.; Masters, D.; Speagle, J.S. The Impossibly Early Galaxy Problem. ApJ 2016, 824, 21, [arXiv:astro-
ph.GA/1506.01377]. https://doi.org/10.3847/0004-637X/824/1/21.
66. Franck, J.R.; McGaugh, S.S. Spitzer’s View of the Candidate Cluster and Protocluster Catalog (CCPC). ApJ 2017, 836, 136,
[arXiv:astro-ph.GA/1701.05560]. https://doi.org/10.3847/1538-4357/836/1/136.
13 of 13

67. Nanayakkara, T.; Glazebrook, K.; Jacobs, C.; Schreiber, C.; Brammer, G.; Esdaile, J.; Kacprzak, G.G.; Labbe, I.; Lagos, C.; Marchesini,
D.; et al. A population of faint, old, and massive quiescent galaxies at 3 < z < 4 revealed by JWST NIRSpec Spectroscopy. arXiv
e-prints 2022, p. arXiv:2212.11638, [arXiv:astro-ph.GA/2212.11638]. https://doi.org/10.48550/arXiv.2212.11638.
68. Glazebrook, K.; Nanayakkara, T.; Schreiber, C.; Lagos, C.; Kawinwanichakij, L.; Jacobs, C.; Chittenden, H.; Brammer, G.; Kacprzak,
G.G.; Labbe, I.; et al. An extraordinarily massive galaxy that formed its stars at zrsim11. arXiv e-prints 2023, p. arXiv:2308.05606,
[arXiv:astro-ph.GA/2308.05606]. https://doi.org/10.48550/arXiv.2308.05606.
69. Sanders, R.H. Cosmology with modified Newtonian dynamics (MOND). MNRAS 1998, 296, 1009–1018, [arXiv:astro-ph/astro-
ph/9710335]. https://doi.org/10.1046/j.1365-8711.1998.01459.x.
70. McGaugh, S.S. Predictions for the Sky-Averaged Depth of the 21 cm Absorption Signal at High Redshift in Cosmologies with
and without Nonbaryonic Cold Dark Matter. PRL 2018, 121, 081305, [arXiv:astro-ph.CO/1808.02532]. https://doi.org/10.1103/
PhysRevLett.121.081305.
71. Gupta, R.P. JWST early Universe observations and ΛCDM cosmology. Monthly Notices of the Royal Astronomical Society 2023,
524, 3385–3395, [https://academic.oup.com/mnras/article-pdf/524/3/3385/50949028/stad2032.pdf]. https://doi.org/10.1093/
mnras/stad2032.
72. Boylan-Kolchin, M. Stress testing ΛCDM with high-redshift galaxy candidates. Nature Astronomy 2023, 7, 731–735, [arXiv:astro-
ph.CO/2208.01611]. https://doi.org/10.1038/s41550-023-01937-7.
73. Haslbauer, M.; Kroupa, P.; Zonoozi, A.H.; Haghi, H. Has JWST Already Falsified Dark-matter-driven Galaxy Formation? ApJ
2022, 939, L31, [arXiv:astro-ph.GA/2210.14915]. https://doi.org/10.3847/2041-8213/ac9a50.
74. Labbé, I.; van Dokkum, P.; Nelson, E.; Bezanson, R.; Suess, K.A.; Leja, J.; Brammer, G.; Whitaker, K.; Mathews, E.; Stefanon,
M.; et al. A population of red candidate massive galaxies 600 Myr after the Big Bang. Nature 2023, 616, 266–269, [arXiv:astro-
ph.GA/2207.12446]. https://doi.org/10.1038/s41586-023-05786-2.
75. Naidu, R.P.; Oesch, P.A.; van Dokkum, P.; Nelson, E.J.; Suess, K.A.; Brammer, G.; Whitaker, K.E.; Illingworth, G.; Bouwens,
R.; Tacchella, S.; et al. Two Remarkably Luminous Galaxy Candidates at z ≈ 10-12 Revealed by JWST. ApJ 2022, 940, L14,
[arXiv:astro-ph.GA/2207.09434]. https://doi.org/10.3847/2041-8213/ac9b22.
76. Finkelstein, S.L.; Bagley, M.B.; Haro, P.A.; Dickinson, M.; Ferguson, H.C.; Kartaltepe, J.S.; Papovich, C.; Burgarella, D.; Kocevski,
D.D.; Huertas-Company, M.; et al. A Long Time Ago in a Galaxy Far, Far Away: A Candidate z 12 Galaxy in Early JWST CEERS
Imaging. ApJ 2022, 940, L55, [arXiv:astro-ph.GA/2207.12474]. https://doi.org/10.3847/2041-8213/ac966e.
77. Atek, H.; Shuntov, M.; Furtak, L.J.; Richard, J.; Kneib, J.P.; Mahler, G.; Zitrin, A.; McCracken, H.J.; Charlot, S.; Chevallard, J.;
et al. Revealing galaxy candidates out to z = 16 with JWST observations of the lensing cluster SMACS0723. MNRAS 2023,
519, 1201–1220, [arXiv:astro-ph.GA/2207.12338]. https://doi.org/10.1093/mnras/stac3144.
78. Adams, N.J.; Conselice, C.J.; Ferreira, L.; Austin, D.; Trussler, J.A.A.; Juodžbalis, I.; Wilkins, S.M.; Caruana, J.; Dayal, P.; Verma, A.;
et al. Discovery and properties of ultra-high redshift galaxies (9 < z < 12) in the JWST ERO SMACS 0723 Field. MNRAS 2023,
518, 4755–4766, [arXiv:astro-ph.GA/2207.11217]. https://doi.org/10.1093/mnras/stac3347.
79. Casey, C.M.; Akins, H.B.; Shuntov, M.; Ilbert, O.; Paquereau, L.; Franco, M.; Hayward, C.C.; Finkelstein, S.L.; Boylan-Kolchin, M.;
Robertson, B.E.; et al. COSMOS-Web: Intrinsically Luminous zrsim10 Galaxy Candidates Test Early Stellar Mass Assembly. arXiv
e-prints 2023, p. arXiv:2308.10932, [arXiv:astro-ph.GA/2308.10932]. https://doi.org/10.48550/arXiv.2308.10932.
80. Yung, L.Y.A.; Somerville, R.S.; Finkelstein, S.L.; Wilkins, S.M.; Gardner, J.P. Are the ultra-high-redshift galaxies at z > 10
surprising in the context of standard galaxy formation models? MNRAS 2023, [arXiv:astro-ph.GA/2304.04348]. https:
//doi.org/10.1093/mnras/stad3484.
81. Harikane, Y.; Ouchi, M.; Oguri, M.; Ono, Y.; Nakajima, K.; Isobe, Y.; Umeda, H.; Mawatari, K.; Zhang, Y. A Comprehensive Study
of Galaxies at z 9-16 Found in the Early JWST Data: Ultraviolet Luminosity Functions and Cosmic Star Formation History at the
Pre-reionization Epoch. ApJS 2023, 265, 5, [arXiv:astro-ph.GA/2208.01612]. https://doi.org/10.3847/1538-4365/acaaa9.
82. Finkelstein, S.L.; Leung, G.C.K.; Bagley, M.B.; Dickinson, M.; Ferguson, H.C.; Papovich, C.; Akins, H.B.; Arrabal Haro, P.;
Dave, R.; Dekel, A.; et al. The Complete CEERS Early Universe Galaxy Sample: A Surprisingly Slow Evolution of the
Space Density of Bright Galaxies at z ~8.5-14.5. arXiv e-prints 2023, p. arXiv:2311.04279, [arXiv:astro-ph.GA/2311.04279].
https://doi.org/10.48550/arXiv.2311.04279.
83. Robertson, B.; Johnson, B.D.; Tacchella, S.; Eisenstein, D.J.; Hainline, K.; Arribas, S.; Baker, W.M.; Bunker, A.J.; Carniani, S.;
Carreira, C.; et al. Earliest Galaxies in the JADES Origins Field: Luminosity Function and Cosmic Star-Formation Rate Density
300 Myr after the Big Bang. arXiv e-prints 2023, p. arXiv:2312.10033, [arXiv:astro-ph.GA/2312.10033]. https://doi.org/10.48550
/arXiv.2312.10033.
84. Melia, F. The cosmic timeline implied by the JWST high-redshift galaxies. MNRAS 2023, 521, L85–L89, [arXiv:astro-
ph.CO/2302.10103]. https://doi.org/10.1093/mnrasl/slad025.
85. McGaugh, S.S. Early Galaxy Formation and the Hubble Constant Tension. Research Notes of the American Astronomical Society 2023,
7, 20. https://doi.org/10.3847/2515-5172/acba9a.
86. Planck Collaboration.; Aghanim, N.; Akrami, Y.; Ashdown, M.; Aumont, J.; Baccigalupi, C.; Ballardini, M.; Banday, A.J.;
Barreiro, R.B.; Bartolo, N.; et al. Planck 2018 results. V. CMB power spectra and likelihoods. A&A 2020, 641, A5, [arXiv:astro-
ph.CO/1907.12875]. https://doi.org/10.1051/0004-6361/201936386.
87. Lewis, A.; Challinor, A. CAMB: Code for Anisotropies in the Microwave Background. Astrophysics Source Code Library, record
ascl:1102.026, 2011, [1102.026].

You might also like